Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Current Pharmaceutical Design, 2007, 13, 2057-2073 2057

Heparanase: Structure, Biological Functions, and Inhibition by Heparin-Derived


Mimetics of Heparan Sulfate
Israel Vlodavsky1,*, Neta Ilan1, Annamaria Naggi2 and Benito Casu2,*

1
Cancer and Vascular and Biology Research Center, The Bruce Rappaport Faculty of Medicine, Technion, Haifa 31096 and 2G. Ron-
zoni Institute for Chemical and Biochemical Research, via G. Colombo, 81, 20133 Milan, Italy

Abstract: Heparanase is an endoglycosidase which cleaves heparan sulfate (HS) and hence participates in degradation and remodeling of
the extracellular matrix (ECM). Heparanase is preferentially expressed in human tumors and its over-expression in tumor cells confers an
invasive phenotype in experimental animals. The enzyme also releases angiogenic factors from the ECM and thereby induces an angio-
genic response in vivo. Heparanase upregulation correlates with increased tumor vascularity and poor postoperative survival of cancer
patients. Heparanase is synthesized as a 65 kDa inactive precursor that undergoes proteolytic cleavage, yielding 8 kDa and 50 kDa pro-
tein subunits that heterodimerize to form an active enzyme. Heparanase exhibits also non-enzymatic activities, independent of its in-
volvement in ECM degradation. Among these, are the enhancement of Akt signaling, stimulation of PI3K- and p38-dependent endothe-
lial cell migration, and up regulation of VEGF, all contributing to its potent pro-angiogenic activity. Studies on relationships between
structure and heparanase inhibition activity of nonanticogulant heparins systematically differing in their O-sulfation patterns, degrees of
N-acetylation, and glycol-splitting of both pre-existing nonsulfated uronic acid residues (prevalently D-glucuronic) and/or those (L-
iduronic acid/L-galacturonic acid) generated by graded 2-O-desulfation, have permitted to select effective inhibitors of the enzymatic ac-
tivity of heparanase. N-acetylated, glycol-split heparins emerged as especially strong inhibitors of heparanase, exerting little or no release
of growth factors from ECM. N-acetylated glycol-split species of heparin, as well as heparanase gene silencing inhibit tumor metastasis,
angiogenesis and inflammation in experimental animal models. These observations and the unexpected identification of a single func-
tional heparanase, suggest that the enzyme is a promising target for anti-cancer and anti-inflammatory drug development.
Key Words: Heparanase, angiogenesis, heparanase inhibitors, heparin, heparin derivatives.

1. INTRODUCTION of modulating the activity of growth factors such as FGF-2 and


Heparan sulfate proteoglycans (HSPGs) are ubiquitous macro- VEGF and enzymes such as thrombin and lipoprotein lipase. The
molecules associated with the cell surface and ECM of a wide range ECM provides an essential physical barrier between cells and tis-
of cells of vertebrate and invertebrate tissues [1, 2]. The basic sues, as well as a scaffold for cell growth, migration, differentiation
HSPG structure consists of a protein core to which several linear and survival, and undergoes continuous remodeling during devel-
heparan sulfate (HS) chains are covalently O-linked. The polysac- opment and in certain pathological conditions such as wound heal-
charide chains are typically composed of repeating hexuronic acid ing and cancer [10]. ECM-remodeling enzymes are thus expected to
and D-glucosamine disaccharide units that are modified at various profoundly affect cell and tissue function. While intensive research
positions by sulfation, epimerization and N-acetylation, yielding focused on enzymes capable of degrading and remodeling protein
clusters of sulfated disaccharides separated by low or non-sulfated components in the ECM [11,12], less attention was paid to enzymes
regions [1-3]. Heparan sulfate (HS) glycosaminoglycans bind to (e.g., heparanase) cleaving glycosaminoglycan side chains.
and assemble ECM proteins (i.e., laminin, fibronectin, collagen Heparanase is an endo--D-glucuronidase capable of cleaving
type IV) and thereby contribute significantly to the ECM' self as- HS side chains at a limited number of intrachain sites, yielding HS
sembly and integrity. HS also play important roles in cell-cell and fragments of still appreciable size (~5-7 kDa) [13-18]. Heparanase
cell-ECM interactions. Moreover, the HS chains, unique in their activity has long been detected in a number of cell types and tis-
ability to bind a multitude of proteins, ensure that a wide variety of sues. Importantly, heparanase activity correlated with the metastatic
bioactive molecules bind to the cell surface and ECM and thereby potential of tumor-derived cells, attributed to enhanced cell dis-
function in the control of normal and pathological processes, among semination as a consequence of HS cleavage and remodeling of the
which are morphogenesis, tissue repair, inflammation, vasculariza- ECM barrier [19, 20]. Similarly, heparanase activity is implicated
tion, and cancer metastasis [1-5]. Apart from sequestration of bioac- in neovascularization, inflammation and autoimmunity, involving
tive molecules, transmembrane (syndecans) and phospholipid- migration of vascular endothelial cells and activated cells of the
anchored (glypicans) HSPGs have a co-receptor role in which the immune system [19-21]. In spite of the attractive clinical relevance
proteoglycan, in concert with the other cell surface molecules, of the pro-metastatic, pro-inflammatory and pro-angiogenic activi-
comprises a functional receptor complex that binds the ligand and ties of heparanase, progress in the field was slow, largely due to the
mediates its action [1-5]. HS also tether a multitude of growth fac- lack of recombinant heparanase, molecular probes, antibodies, and
tors, chemokines, cytokines and enzymes to the ECM and cell sur- a simple bioassay to quantify heparanase activity.
face, providing a low affinity storage depot for heparin-binding Heparanase activity was attributed to proteins with molecular
proteins [4], including the potent pro-angiogenic factors FGF-2 and weight ranging from 8 to 130 kDa, raising the possible existence of
VEGF [6-9]. Cleavage of HS side chains is therefore expected not several HS-degrading endoglycosidic enzymes [19-21]. This confu-
only to alter the integrity of the ECM, but also to release HS-bound sion was solved when the cloning of a single human heparanase
biological mediators. In addition, HS fragments are also capable cDNA sequence was independently reported by several groups [22-
25]. With the availability of appropriate reagents, heparanase re-
search entered a new era. Recent studies have shown that hepara-
*Address correspondence to these authors at the Vascular and Tumor Biol- nase is up regulated in an increasing number of primary human
ogy Research Center, Rappaport Faculty of Medicine, Technion, P.O. Box
9649, Haifa 31096, Israel; Tel: 972-4-8295410; Fax: 972-4-853947;
tumors [26]. Heparanase upregulation correlates with increased
E-mail: Vlodavsk@cc.huji.ac.il lymph node and distant metastasis, increased micro-vessel density,
G. Ronzoni Institute for Chemical and Biochemical Research – via G. Co- and reduced post-operation survival of cancer patients [19,20,26,
lombo, 81, Milan 20133, Italy; Tel: +39-02-7064-1623; Fax: +39-02- 7064- 27], providing a strong clinical support for its pro-metastatic and
1634; E-mail: casu@ronzoni.it pro-angiogenic features and for heparanase as a promising target for

1381-6128/07 $50.00+.00 © 2007 Bentham Science Publishers Ltd.


2058 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

the development of anti-cancer treatments. Here, we summarize erodimer composed of the 8 and 50 kDa subunits. Indeed, attempts
recent progress in molecular and cellular aspects of heparanase, and to express the truncated 50 kDa (Lys158-Ile543) protein alone yielded
outline strategies for the development of heparanase inhibitors. no enzymatic activity [39], indicating that an N-terminus sequence
Taking into account the estabished role of HS in cancer [28,29] and is required. This hypothesis was confirmed by co-transfection and
recent clinical trials demonstrating a beneficial effect of heparin in immunoprecipitation approaches, convincingly demonstrating that
cancer patients [30-33], we discuss relationships between structure the two subunits are associated with each other, and that enzymatic
and heparanase inhibition activity of heparin derivatives, with spe- activity is only obtained by co-expression of both the 8 and 50 kDa
cial emphasis on non-anticoagulant glycol-split species which spe- subunits [36,37]. Multiple sequence alignment and secondary struc-
cifically inhibit heparanase enzymatic activity and experimental ture prediction suggest that heparanase adopts a TIM barrel fold,
metastasis. similar to other glycosyl hydrolases [39]. This fold motif usually
consists of eight alternating -helices and -strands. Within the
2. MOLECULAR AND BIOCHEMICAL PROPERTIES OF heparanase 50 kDa subunit, clear homology was noted for only six
HEPARANASE / units, leading Nardella et al. to suggest that the two other units
2.1. Structure are contributed by the 8 kDa subunit. Indeed, structural prediction
Only one gene (HPSE) has been shown to encode for a protein revealed the presence of a // element in the 8 kDa subunit,
with heparanase activity [22-25]. Sequence analysis revealed that which may thus contribute the missing TIM barrel units [45]. These
heparanase is highly conserved, with similar sequences found in and other [37] studies, indicated that the linker domain (Ser110-
human, rat, mouse, cow, chicken, molluscs and zebra fish [19, 34]. Gln157) inhibits heparanase activity and needs to be fully removed.
The human heparanase cDNA contains an open reading frame that Adopting site-directed mutagenesis to identify amino acids essential
encodes a polypeptide of 543 amino acids with a molecular weight for cleavage at Glu109-Ser110 (site 1) and Gln157-Lys158 (site 2), Ab-
of 61.2 kDa. The active heparanase purified from placenta, plate- boud-Jarrous et al. reported that none of the mutations generated at
lets, and various cell lines was found to lack its N-terminal 156 site 1 and its flanking regions had an effect on heparanase process-
amino acids, suggesting post-translational proteolysis of the ing and activity [46]. In contrast, substitution of Tyr156 (site 2) by
heparanase polypeptide [22-25]. In fact, active heparanase was alanine or glutamine rendered heparanase inactive and improperly
subsequently reported to be a heterodimer consisting of a 50 kDa processed [46]. Subsequent studies revealed that a bulky hydropho-
subunit (Lys158–Ile543) associated non-covalently with an 8 kDa bic amino acid (i.e., Tyr156) at position 2 (P2) of the cleavage site
peptide (Gln36–Glu109). The intervening 6 kDa peptide (Ser110– (Gln157-Lys158) is absolutely required for heparanase processing and
Gln157) is excised by proteolysis [35-37]. Based on the predicted activation, resembling the cleavage specificity of cathepsin L [46].
amino acid sequence, the 50 kDa subunit of human heparanase Indeed, incubation of purified latent 65 kDa heparanase with
contains six putative N-glycosylation sites. Although glycosylation cathepsin L yielded properly processed and active heparanase,
was not required for enzyme activity, secretion of heparanase was composed of the 50 and 8 kDa subunits. Processing and activation
regulated by glycosylation [38]. The sequence also contains a 35- of the pro-enzyme by intact cells and in a cell free system was in-
amino acid N-terminal signal sequence (Met1–Ala35) and a C- hibited in the presence of a specific, cell permeable inhibitor of
terminal hydrophobic domain (Pro515-Ile534). Protein sequence cathepsin L [46]. Applying a structural model, we demonstrated
alignment approaches in combination with secondary structure that the linker segment, or even a small 1 kDa portion at its C-
predictions indicated that heparanase contains sequences that are terminus, render the active site inaccessible to the HS substrate.
homologous to families 10, 39, and 51 of the clan A glycosyl hy- Moreover, heparanase activity was inhibited in the presence of a 1
drolyses, especially in terms of the active-site regions [39]. This kDa peptide corresponding to the C-terminus of the linker segment.
clan of enzymes uses a general acid catalysis mechanism for the While predicted model structures do provide important information
hydrolysis of glycosidic bonds. The mechanism requires two criti- (Fig. 1), it is limited by the relatively low sequence homology with
cal residues, a proton donor and a nucleophile, both of which ap- other glycosyl hydrolases, and awaits further confirmation by crys-
pear to be conserved in heparanase at Glu225 and Glu343, respec- tallization and x-ray analysis. Site directed mutagenesis of aromatic
tively [39]. Site-directed mutagenesis of these residues completely residues along the linker segment, supports a mechanism by which
abolished heparanase activity, indicating that heparanase uses a cathepsin L is not only responsible for cleavage at site 2, but is
catalytic mechanism characteristic of clan A glycosyl hydrolases actually capable of a stepwise cleavage and removal of the entire
[39]. linker region, suggesting that processing and activation of pro-
heparanase can be brought about solely by cathepsin L and/or
2.2. Processing and Activation cathepsin L-like proteases (our unpublished results). It is likely that
cathepsins other than cathepsin L may activate pro-heparanase,
Taking into account the multitude of polypeptides associated
possibly in a cell- and tissue- dependent manner.
with HS on the cell surface and ECM and their ability to profoundly
affect cell and tissue function, heparanase activity and bioavail- Recent findings indicated that heparanase processing occurs in-
ability should be kept tightly regulated. While regulation at the tracellularly [47-49], pointing to acidic vesicles, most likely
transcriptional level [40-44] represents one type of control mecha- lysosomes, as the processing organelles. Following exogenous addi-
nism, regulation at the post-translational level (i.e., heparanase tion, heparanase was noted to reside within perinuclear endocytic
processing, cellular localization and secretion) provides an addi- vesicles identified as late endosomes [50] and lysosomes [51]. Ap-
tional key regulatory mechanism. A major 50 kDa protein is de- plying anti-heparanase antibodies that distinguish between the la-
tected in cell lysates following transfection and over expression of tent and processed heparanase forms, we have demonstrated that
the heparanase cDNA, correlating with high levels of enzymatic not only the processed, but also the 65 kDa latent heparanase was
activity [25]. In contrast, a 65 kDa protein was found in the cell localized in endocytic vesicles, indicating that processing does not
conditioned medium, raising the possibility that the protein is first take place at the cell surface [49]. Complete inhibition of hepara-
synthesized as a latent proenzyme, which is then activated by pro- nase processing by chloroquine and bafilomycin A1, reagents that
teolytic processing. Purification of the human heparanase to homo- raise the pH of acidic vesicles and thus inhibit the enzymatic activ-
geneity enabled N-terminus sequencing of the 50 kDa protein [35]. ity of resident enzymes, further points to acidic vesicles as the
Interestingly, the purified heparanase preparation was noted to in- heparanase processing site [49]. Taking this notion a step further,
clude an 8 kDa protein and further analysis revealed that this pro- Cohen et al. utilized a cell fractionation approach and demonstrated
tein is derived from the N-terminus region (Gln36-Glu109) of that lysosomal/endosomal, but not cytoplasmic preparation is capa-
heparanase, immediately next to the signal sequence [35]. This ble of heparanase processing, yielding an active enzyme [52].
finding led the authors to suggest that active heparanase is a het- Moreover, processing by the lysosomal/endosomal preparation was
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2059

Formula
CH2OSO3- CO2- CH2OSO3-
O O O
OH O OH(SO3)
1 (AGA) O OH O

NHSO3- OH NHSO3-
(Ac)

CH2OSO3- CH2OSO3- COO-


O O O O
2 (H) OH
CO2- O OH OH
O
OH O
O
n>>m
NHSO3- OSO3- n NHAc OH m
(OH)

GlcNSO36SO3 - IdoA2SO3 GlcNAc6SO3 - GlcA

CH2OSO3- CH2OSO3- COO-


O O O O
CO2- O O
3 (NAH) OH
OH
OH
O
OH
O
n>>m NHAc OSO3- n NHAc OH m
(OH)

GlcNAc6SO3 - IdoA2SO3 GlcNAc - GlcA

CH2OSO3- CH2OSO3-
O O O O
50 CO2- COO- O
4 ( H gs ) OSO3-
OH
O OH
O
O

NHR OSO3- NHR OH OH n

CH2OSO3- CO2- CH2OSO3- CH2OSO3-


O O O O O
CO2-
5 (AGA*IA) OH
OH
O OSO3-
OH
O OH
O O O

NHAc OH NHSO3- OSO3- NHSO3-

CH2OSO3- CO2- CH2OSO3- CH2OSO3-


O O O O O
CO2-
OH O OSO3- O OH
6 (gs-AGA*IA) O O OH O

NHAc OH OH NHSO3- OSO3- NHSO3-

CH2OSO3- CH2OSO3- CO2-


O O O O O
CO - CO2-
7 aza-oligoH OH2
O
OH
O OH
O OH
O OH
NR

OSO3- OSO3 OSO3-


OSO3- OSO3-

CH2OSO3- CH2OSO3- CH2OSO3- OSO3- CH2OSO3- CH2OSO3-


O O O O O O
8 MHS OSO3- OSO3- OSO3- OSO3- OSO3- OSO3-
O O O OSO3- O O
OSO3-
OSO3- OSO3- OSO3- OSO3- OSO3- OSO3-
2060 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

Fig. (1). Primary structure, critical amino acids and predicted three-dimensional structure of the heparanase heterodimer. Pre-proheparanase harbors 35 amino
acids signal peptide (s.p., Met1 -Ala35) which is removed upon entering the ER. The protein is then subjected to glycosylated at six N-glycosylation sites ( )
and secreted as a latent ~65 kDa protein (upper panel). Proteolytic processing removes the linker domain (Ser110–Gln157), resulting in an 8 kDa (Gln36–Glu109)
and 50 kDa (Lys158–Ile543) protein subunits (middle panel) that heterodimerize to yield an active enzyme. A predicted three-dimensional structure of the hepara-
nase heterodimer (bottom panel was created based on homology with 1,4--xylanase. Shown (left) are the 8 (yellow) and 50 kDa (gray) subunits and glutamic
acid residues 225 and 343 that comprise the enzyme active site (red). The heparin binding domains (HBD 1 and 2, blue and green) are in close proximity to the
enzyme active site (red). [See text.]

most efficient at acidic pH conditions (pH 4-5) [52], typical of the had no effect on heparanase secretion from tumor-derived cells (our
lysosomal compartment. These results and the ability of cathepsin unpublished results), suggesting that effective stimuli may vary
L, a characteristic lysosomal enzyme, to properly process and acti- among cell types and biological settings. Our recent studies indicate
vate pro-heparanase, strongly support the lysosomal compartment that nucleotides, such as ATP, ADP and adenosine are highly effec-
as the site of processing. tive in stimulating secretion of both active heparanase and cathepsin
D by tumor cells via activation of the PKC and PKA pathways
2.3. Secretion [Shafat et al. J Biol Chem 2006; 281: 23804-11].
In spite of its localization to a highly active protein degradation
environment such as the lysosome, heparanase exhibits a half life of 2.4. Cellular Uptake and Extracellular Retention
about 30 hours [47], relatively long compared with a t1/2 of 2-6 Apart from storage in the endosomal/lysosomal compartment,
hours, and 25 minutes of transmembrane and GPI-anchored efficient uptake of exogenous heparanase by primary fibroblasts
HSPGs, respectively. Residence and accumulation of heparanase in and EC, as well as by tumor-derived cells [47, 49, 50] provides an
late endosomes and lysosomes may indicate that the enzyme func- additional mechanism that limits retention of the enzyme extracel-
tions in physiological turnover of cellular HSPGs [53]. Being the lularly. Several lines of evidence indicate that heparanase uptake is
enzyme not readily accessible to its extracellular substrate, this mediated by cell surface HS. We have demonstrated that addition of
suggests the existence of regulatory mechanism(s) by which intra- heparin or xylosides results in accumulation of heparanase in the
cellular, lysosomal heparanase is secreted in response to local or culture medium of heparanase transfected cells. Heparanase uptake
systemic cues. Recent observations may support the occurrence of was attenuated in HS-deficient cells and in cells that were treated
such a scenario. For example, treatment of human microvascular with bacterial heparinase, but not with chondroitinase ABC. In
endothelial cells (EC) with the pro-inflammatory cytokines TNF addition, transfection and over expression of heparanase in HS-
and IL-1 resulted in a marked increase of heparanase secretion deficient cells resulted in accumulation of the latent pro-enzyme in
[54]. Secretion of heparanase in response to TNF was also noted the culture medium, concomitant with decreased levels of the intra-
in human peripheral T-cells [55]. Interestingly, TNF and IL-1 cellular processed enzyme [47]. This result suggests that intracellu-
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2061

lar accumulation of processed heparanase occurs following uptake wound granulation tissue and blood vessels [61]. Heparanase con-
of the secreted latent protein [47] (Fig. 2). Sequence alignment of tribution to wound healing and wound angiogenesis has been dem-
heparin binding domains (HBD) in the heparanase molecule re- onstrated in several experimental settings. Increased amounts of
vealed the existence of two domains that match consensus se- heparanase were found in the wound fluid of heparanase trans-genic
quences for heparin binding. These were mapped at Lys158-Asp162 (hpa-tg) vs. control mice [61], in agreement with heparanase ex-
at the N-terminus of the 50 kDa heparanase subunit, and at Pro271- pression in healing wounds. Moreover, elevated heparanase levels
Met278 [56]. A peptide containing the Lys158-Asp162 sequence in the wound fluid correlated with a comparable elevation of FGF-2
(KKDC) exhibited firm binding to heparin and HS, and inhibited [61], providing an in vivo support for the ability of heparanase to
both heparanase uptake and enzymatic activity, most likely due to release heparin-binding pro-angiogenic factors. Enhanced and per-
competition with the HS substrate [56]. Furthermore, heparanase sistent wound angiogenesis was further demonstrated in the hpa-tg
deletion mutants lacking each of the heparin binding domains ex- mice by applying magnetic resonance imaging (MRI) [61]. This
hibited no enzymatic activity. Deletion of the KKDC sequence effect was inhibited by a non-anticoagulant glycol-split heparin
(6515) resulted in intracellular accumulation of the 65 kDa pro- (ST1514, corresponding to H50gs in Ref. [62], formula 4), endowed
enzyme that failed to get secreted. Deletion of the Pro271-Met278 with a heparanase inhibitory activity at nM concentrations [62] (see
sequence (6510) led to accumulation of the pro-enzyme in the cell Section 4.2), thus clearly supporting a role for heparanase in wound
conditioned medium [56], further supporting a critical role for HS angiogenesis [61]. In a rat/flap punch model, topical application of
in heparanase uptake and processing. 2D NMR analysis revealed a highly active recombinant heparanase improved wound healing by
strong interaction of the two putative HS binding regions of 40% and enhanced wound angiogenesis [61]. Measurements taken
heparanase with a heparin/HS pentasaccharide (i.e., AGA*IA, for- in the area of flap incisions revealed a significant increase in epithe-
mula 5) [56]. Notably, the KKDC sequence, and specifically ly- lium thickness, suggesting that heparanase promotes keratinocyte
sine158 and lysine159 involved in interaction of the enzyme with proliferation due to an improved bioavailability of factors such as
heparin/HS [56], appeared to reside in close proximity to Glu225 and KGF and HB-EGF. In addition, immunostaining of wound sections
Glu343, comprising the enzyme' active site in a micro pocket domain with anti-smooth muscle actin (SMA) antibody revealed a seven
[39] (Fig. 1). The predicted 3D model further emphasizes the micro fold increase in SMA-positive blood vessels in response to hepara-
pocket region and, in particular, the KKDC sequence (Fig. 1) as a nase treatment [61]. Thus, heparanase accelerates wound healing by
valid target for the development of heparanase inhibiting molecules. enhanced migration and proliferation of keratinocytes and stimula-
More recently, Vreys et al. have identified two additional cell sur- tion of wound blood vessel formation and maturation. The coordi-
face receptors that mediate heparanase uptake, namely the low den- nated, simultaneous release of a combination of HS-bound growth
sity lipoprotein receptor-related protein (LRP) and the mannose 6- factors (i.e., FGF-2, VEGF, HB-EGF, KGF) is unique to hepara-
phosphate receptor [48]. The precise contribution of each receptor nase and may account for its efficient neovascularization and
species to heparanase uptake is yet to be demonstrated. Collec- wound healing promoting effect. It should be noted that wound
tively, the above described studies emphasize the complexity and healing is only one example for the involvement of heparanase in
tight regulation of heparanase expression, processing and secretion tissue remodeling and neovasculari-zation. For example, hepara-
(Fig. 2) supporting its potency and significance in normal and nase transgenic mice (hpa-tg) exhibit excess branching and widen-
pathological conditions. ing of mammary gland ducts, and an accelerated rate of hair
growth, both accompanied by enhanced vascularization [63, 64].
3. HEPARANASE AND CANCER PROGRESSION
3.1. Pro-Angiogenic Properties 3.2. Non-Enzymatic Functions Associated with Angiogenesis,
Cell Survival and Migration
HSPGs are prominent components of blood vessels, and HSPG
degrading enzymes have long been implicated in a number of angi- Akt Activation
ogenesis-related cellular processes. A critical early event in the Heparanase over expression in human U87 glioma [65], HT 29
angiogenic process is degradation of the subendothelial basement colon carcinoma [66], CAG myloma [67], MCF7 [68], MDA-MB-
membrane (BM), followed by endothelial cell (EC) migration to- 231 [69], and MDA-MB-435 [70] breast carcinoma cells correlated
ward the angiogenic stimulus. Similar to its involvement in tumor with enhanced xenograft tumor growth and vascularization. Using
cell dissemination, it is conceivable that by degrading HS in the the RIP-Tag2 tumor model, Joyce et al. have recently demonstrated
BM, heparanase may directly facilitate EC invasion and sprouting. elevated levels of heparanase mRNA and protein upon the transi-
Indeed, heparanase expression by FGF-2 stimulated bone marrow- tion from normal to angiogenic islets, which further increased when
derived EC was demonstrated by RT-PCR [57]. Immunohisto- solid tumors were detected [71]. These studies support the notion
chemistry of tumor specimens revealed heparanase staining of EC that heparanase not only facilitates tumor metastasis, but also con-
in capillaries, but not mature blood vessels [57,58]. Moreover, by tributes to the angiogenic switch and subsequent growth of the pri-
releasing HS-bound angiogenic growth factors (i.e., FGF-2, VEGF) mary tumor. Enhanced tumor progression correlated with elevation
from the ECM [7] active heparanase may indirectly facilitate EC in blood vessel density, revealed by staining with anti-PECAM-1
migration and proliferation [59]. In fact, given the multitude of antibodies, as well as by MRI analysis [68]. At the molecular level,
biological mediators that are sequestered by HS in the ECM [60], heparanase over expression was noted to facilitate cell adhesion and
heparanase activity liberates a number of active molecules that may migration of tumor cells, primary EC and T lymphocytes, mediated,
act cooperatively or synergistically to promote neovascularization. al least in part, by 1-integrin and Rac activation [55, 65, 72].
An idealized picture of heparanase-induced release and activation Heparanase over expression in U87 glioma, as well as in several
of FGF-2 molecules originally sequestered by HS chains of HSPGs other tumor derived cells, correlated with enhanced serine/threonine
is shown in Fig. (3). protein kinase Akt/PKB phosphorylation levels [65]. Moreover,
Wound healing orchestrates multiple cell types (i.e., neutro- exogenous addition of heparanase to primary EC markedly stimu-
phils, macrophages, fibroblasts, keratinocytes, endothelial cells), lated Akt phosphorylation [73]. This effect occurred both in the
soluble (i.e., growth factors, cytokines, chemokines) and insoluble presence or absence of HS on the cell surface, was independent of
(ECM components) mediators in a complex sequence of events. heparanase enzymatic activity, and was augmented by heparin [73].
Orchestration and regulation of the rapidly developing new tissue At the cellular level, heparanase addition stimulated PI 3-kinase-
observed in wound healing depend not only on cells and bioactive dependent EC migration and invasion, and significantly improved
polypeptides, but also on the ECM microenvironment, and require EC rearrangement into lumen containing tube-like structures [73].
new blood vessel formation to nourish the newly formed granula- These observations imply that heparanase is capable of eliciting
tion tissue. Elevated heparanase expression was observed in the angiogenic responses by a direct effect on EC. The ability of
2062 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

Fig. (2). A schematic presentation of a proposed model for heparanase biosynthesis, processing and trafficking. Pre-pro-heparanase is first targeted to the ER
lumen via its own signal peptide (Met1-Ala35, 1). The 65 kDa pro-heparanase is then shuttled to the Golgi apparatus, and is subsequently secreted via vesicles
that bud from the Golgi (2), a step that is specifically inhibited by Brefeldin A (BFA) [50]. Once secreted, heparanase rapidly interacts with cell membrane
HSPGs such as syndecan-family members (3) [47], mannose-6 phosphate receptor or LRP [48], followed by a rapid endocytosis of the heparanase-HSPG
complex (4) that appears to accumulate in endosomes [50]. This step is inhibited by cytochalasin D (Cyto. D) [50], heparinase [47], or heparin [47]. Conver-
sion of endosomes to lysosomes results in heparanase processing and activation (5) that, in turn, participates in the turnover of HS side chains in the lysosome.
Heparanase processing and activation is specifically inhibited by chloroquine and Bafilomycin A, inhibitors of lysosomal proteinases. Typically, heparanase
appears at perinuclear lysosomal vesicles (6). Such a trafficking route may be bypassed by several potential ways, such as direct conversion of exocytosed
vesicles to endosomes (dashed arrow). Lysosomal heparanase may translocate to the nucleus, where it affects gene transcription, thus contributing to a more
differentiated state of carcinoma cells [120, 121] (7), or can get secreted in response to local or systemic cues. Secretion of active heparanase heterodimer is
inhibited by PKC inhibitors (Bis), and P2Y receptor antagonists (MRS, PPADS) (8). The latent secreted heparanase can also interact with heparanase-binding
protein (HBP) and activate signaling components such as Akt, p38, Src, Pyk2 and integrins, leading to enhanced cell adhesion, migration, VEGF induction and
angiogenesis [47, 49, 55, 65, 72 ] (9).

heparanase to stimulate Akt suggests that heparanase may protect mately involved in VEGF gene regulation. Interestingly, VEGF
tumor cells from apoptosis [68], although the survival promoting elevation by heparanase correlated with increased p38 phosphoryla-
mechanism has not been sufficiently elucidated. The ability of ex- tion levels, a signaling pathway implicated in VEGF induction [70].
ogenously added heparanase to activate signal transduction cas- Nonetheless, p38 inhibitors had no effect on heparanase-mediated
cades, and its augmentation by heparin may indicate the existence VEGF up regulation, suggesting the operation of another signaling
of a cell surface heparanase receptor (Fig. 2, HBP), possibly low pathway(s) elicited by heparanase. Screening of several additional
density lipoprotein receptor-related proteins (LRP) [48], yet this inhibitors led to the identification of Src, a kinase that was shown to
aspect awaits further research and proper confirmation. modulate VEGF transcription [74-76], as a mediator of VEGF up
VEGF up Regulation regulation by heparanase. Moreover, Src inhibitors prevented
VEGF induction by heparanase and significantly attenuated cell
In addition to the above described pro-angiogenic effects attrib- migration enhanced by heparanase, positioning Src as a critical
uted to heparanase enzymatic and non-enzymatic activities, hepara- down stream component that mediates heparanase functions [70].
nase is also closely involved in VEGF gene regulation. Transfection VEGF induction and Src activation require heparanase secretion
and over expression of heparanase in rat C6 glioma, MDA-MB-435 that was recapitulated by exogenous addition of the enzyme. Impor-
human breast carcinoma and human embryonic kidney HEK293 tantly, Src activation was noted also upon exogenous addition of
cells, were accompanied by a 3-6 fold increase in VEGF mRNA point mutated (Glu225, Glu343) heparanase that lacks enzymatic ac-
and protein levels, correlated with enhanced Matrigel and tumor tivity. Thus, heparanase exerts enzymatic activity-dependent (i.e.,
xenograft vascularization [70]. Moreover, transfection of the highly release of FGF-2) and independent (i.e., VEGF induction) pro-
metastatic mouse B16-BL6 melanoma cells with heparanase- angiogenic effects. It therefore appears that in addition to its pro-
specific siRNA resulted in a 75% decrease in heparanase mRNA metastatic function (see below), heparanase affects several key
and VEGF levels. This implies that endogenous heparanase is inti-
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2063

Fig. (3). Idealized representation of cleavage of HS chains of a heparan sulfate proteoglycan (HSPG) by heparanase, with release of growth factors (GF), their
activation by dimerization favored by HS fragments, and formation of signaling ternary complexes with growth factor receptors (GFR) on the surface of endo-
thelial cells.

components in tumor progression, resulting in increased blood ves- control ribozyme. Similarly, lung colonization of B16-BL6 mela-
sel density and maturation, enhanced tumor cell motility, and acti- noma cells was markedly (> 90%) reduced applying cells trans-
vation of signaling mediators that govern tumor cell proliferation fected with anti-heparanase siRNA due to a marked inhibition of
(i.e., Src) and survival (Akt). Collectively, these effects position both heparanase gene expression and enzymatic activity, naturally
heparanase as an attractive target for the development of anti cancer expressed by these cells [80]. Subcutaneous primary tumors pro-
drugs. duced by hpa-Eb cells expressing the anti-heparanase ribozyme
were less vascularized, supporting the pro-angiogenic function of
3.3. Pro-Metastatic Properties the enzyme. Altogether, both over-expression and silencing of the
Heparanase activity correlated with the metastatic potential of heparanase gene clearly indicate that heparanase not only enhances
mouse B16 melanoma [77,78] and Eb lymphoma [79] cells. Thus, cell dissemination, but also promotes the establishment of a vascu-
sub lines with higher potential for metastasis and organ colonization lar network that accelerates primary tumor growth and provides a
exhibited a higher enzymatic activity than low- or non- metastatic gateway for invading metastatic cells.
cells. These early observations gained substantial support when While the studies described above provide a proof-of-concept
specific molecular probes became available shortly after cloning the for the pro-metastatic and pro-angiogenic capacity of heparanase,
heparanase gene. Hulett et al. employed Northern blot analysis to the clinical significance of the enzyme in tumor progression
study heparanase expression in cells and tissues [22], while Vlo- emerges from a systematic evaluation of heparanase expression in
davsky et al. utilized a transfection approach [25]. In these studies, primary human tumors. Immunohistochemistry, in situ hybridi-
highly metastatic rat mammary adenocarcinoma cell lines were zation, RT-PCR and real time-PCR analyses revealed that hepara-
noted to express high levels of heparanase mRNA transcripts com- nase is up regulated in essentially all human tumors examined [26,
pared with their non metastatic counterpart cells [22]. Subcutaneous 81, 82]. Heparanase up regulation in primary human tumors corre-
inoculation of non metastatic Eb lymphoma cells engineered to over lated in some cases with tumors larger in size and with enhanced
express heparanase (hpa-Eb) resulted in a significant decease in micro vessel density, as well as with reduced post-operative sur-
survival time of the mice due to a massive liver infiltration [25], vival of cancer patients providing a clinical support for the pro-
further supporting the correlation between heparanase expression angiogenic function of the enzyme [19,20,26,27].
and the metastatic capacity of cancer cells.
We employed siRNA and ribozyme technologies to reduce 4. INHIBITION OF HEPARANASE
heparanase expression levels in a specific manner. Transfection and 4.1. Heparanase Inhibition Strategies
stable expression of anti-heparanase ribozyme construct in human
MDA-MB-435 breast carcinoma cells, known to express high levels Attempts to inhibit heparanase enzymatic activity were initiated
of heparanase activity, or in Eb mouse lymphoma cells engineered already at the early days of heparanase research, in parallel with the
to over express the human heparanase gene (hpa-Eb), resulted in a emerging clinical relevance of this activity. More recently, with the
marked decrease in heparanase levels evaluated by RT-PCR and availability of recombinant heparanase and the establishment of
heparanase enzymatic activity. This decrease correlated with 55 - high-throughput screening methods, a variety of inhibitory mole-
65% reduction in cellular invasion through a reconstituted basement cules have been developed, including neutralizing antibodies, pep-
membrane (Matrigel) [80]. Moreover, mice inoculated (s.c) with tides, small molecules, modified non-anticoagulant species of hepa-
hpa-Eb lymphoma cells transfected with anti-heparanase ribozyme rin, as well as several other polyanionic molecules, such as lamina-
exhibited a marked decrease in liver metastasis and survived sig- ran sulfate, suramin and PI-88 [26,27]. Heparanase can be inhibited
nificantly longer than mice inoculated with cells transfected with either at the level of the pro-enzyme by preventing its intracellular
2064 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

processing and activation, or the active enzyme, mostly extracellu- 4.2. Heparin and Heparin Derivatives as Mimetics of Heparan
larly. Heparanase inhibitors may be targeted to the enzyme active sulfate and Heparanase Inhibitors
site, as well as to one or both of its identified heparin/HS binding As mentioned in Section 2, the active form of heparanase is a
regions. Although the present review is especially focused on hepa- 50 kD protein containing a site responsible for cleavage of glycosi-
rin-like, HS mimics, a brief account of alternative reported hepara- dic bonds of -D-glucuronic acid (GlcA) residues of HS chains.
nase-inhibitory strategies is given. The minimum HS motif recognized and cleaved by the the enzyme
Peptides Competing with the Heparin/HS Binding Domains of is trisaccharide 1, where the arrow indicates the site of cleavage.
Heparanase The GlcA residue is flanked by two 6O-,N-sulfated glucosamine
As mentioned in Section 2.4, three potential heparin-binding residues, the first of which could alternatively be N-acetylated and
domains in the heparanase molecule were identified and evidence the second one also 3-O-sulfated [14,16]. Cleavage by heparanase
was provided that one of these domains is mapped at the N- can be observed when sequence 1 is part of pentasaccharides [A.
terminus of the 50 kDa heparanase subunit [56]. A peptide corre- Bisio, J.-P. Li, et al., unpublished; M. Petitou, personal communica-
sponding to this region (KKDC peptide: Lys 158 -Asn 171) physi- tion).
cally associates with heparin and HS and inhibits heparanase enzy- As also described in Section 2, heparanase (both the latent and
matic activity in a dose-responsive manner. Furthermore, antibodies active forms) contains three basic clusters, two of which may be
directed to this region inhibited heparanase activity and a deletion involved in tight binding of HS [56]. Widely pursued strategies for
construct lacking this domain exhibited no enzymatic activity [56]. inhibiting heparanase aim at preventing binding of HS chains to the
We have introduced a cysteine residue at the peptide C-terminus enzyme. Among other strategies (see Section 4.1), this goal could
and found that oxygenation and spontaneous dimeri-zation of the be achieved using HS mimics that compete with HSPGs for the
KKDC peptide significantly improved its heparanase-inhibiting basic clusters of heparanase involved in binding and recognition of
activity. HS chains. Two chemically fully sulfated oligosaccharides (malto-
Studies on processing and activation of the 65 kDa heparanase hexaose sulfate, MHS, and phosphomannopentaose sulfate, PI-88),
proenzyme revealed that the linker peptide must be totally removed were shown to be good heparanase inhibitors [27,86]. Being not
in order to enable interaction between heparanase and its substrate close mimics of the natural substrate of heparanase, MHS and PI-88
[45, and our unpublished results]. As mentioned in section 2.2, a 1 are not covered in detail in the present review.
kDa peptide corresponding to the C-terminus of the linker segment Heparin, an animal polysaccharide belonging to the glycosa-
was found to inhibit heparanase enzymatic activity, most likely by minoglycan (GAG) family, is one of the closest mimics of HS [3]
hindering accessibility of the HS substrate to the enzyme' active and is a natural choice as heparanase inhibitor. Indeed, heparin has
site. been shown to be a potent inhibitor of the enzyme [87, 88]. How-
Small Molecules Targeting at the Catalytic Domain ever, the strong anticoagulant activity of this GAG, mainly associ-
ated with specific pentasaccharide sequences (AGA*IA, formula 5)
Non carbohydrate, small molecule heparanase inhibitors, exem- that constitute the active site for antithrombin (AT) [3], makes
plified by 1H-isoindole-5-carboxylic acid and benzoxazol-5-yl- heparin unsuitable for long-term treatments. As represented in For-
acetic acid, were reported. A novel class of 1-[4-(1H-benzo-imida- mula 2, heparin is prevalently made up by regular disaccharide
zol-2-yl)-phenyl]-3-[4-(1H-benzoimidazol-2-yl)-phenyl]-ureas were sequences of L-iduronic acid 2-sulfate and N-and 6-O-sulfated
recently described as potent inhibitors of heparanase [27]. Among glucosamine, but it is also constituted by undersulfated, GlcA- and
these, compound 1,3-bis-[4-(1H-benzoimidazol-2-yl)-phenyl]-urea GlcNAc-containing sequences, mostly arranged in blocks [3, 89].
displayed a high heparanase inhibitory activity (IC50 - 0.075KM) in Heparin accordingly contains also some of sequences 1 recognized
vitro and a good efficacy in a B16 experimental metastasis model and cleaved by heparanase [14, 16] hence acting as a substrate for
[83]. the enzyme, with behavior as inhibitor conceivably depending on
Monoclonal Antibodies and Microbial Metabolites content and molecular environment of cleavable sequences. An-
An attractive approach for the inhibition of heparanase is the other undesired property of heparin is its capability to displace
development and use of neutralizing monoclonal antibodies to the growth factors sequestered by HS chains in the ECM and on cell
protein. A monoclonal antibody has been reported which effectively surfaces [87]. When complexed with heparin, these growth factors
abolishes the activity of recombinant heparanase when used at a are activated and form with receptors of growth factors (GFR) ter-
protein : antibody ratio of 1:10 [27]. A neutralizing polyclonal anti- nary complexes that trigger mitogenic signals, as illustrated in Fig.
body directed against a peptide corresponding to the heparanase (3) for HS fragments released by heparanase [4]. Ideal hepar-
active site (residues G215 – D234) effectively inhibited restenosis anase inhibitors should accordingly be nonanticoagulant heparin
in a rat model [84]. Random, high-throughput screening of chemi- species endowed with potent antiheparanase activity without favor-
cal libraries and microbial metabolites, and rational design of com- ing release and activation of growth factors.
pounds that block the heparanase active site or ligand-binding do- Early screening of heparin derivatives permitted to identify
main are among the other approaches applied to develop effective some structural features of the polysaccharide associated with inhi-
heparanase inhibitors [26,27,83]. Natural endogenous heparanase bition of heparanase. No significant differences were found be-
inhibitors may also be identified. Further defining the heparanase tween the currently used unfractionated heparins, low-molecular
substrate specificity, catalytic and non-catalytic activities, as well as weight heparins (LMWH) and a tetradecasaccharide heparin frag-
the enzyme X-ray crystal structure is needed for pursuing a more ment; however, a heparin hexasaccharide is a poor inhibitor of the
‘rational’ approach to develop effective and highly specific hepara- enzyme. Typical heparanase-inhibition curves, showing the gel
nase inhibiting molecules. filtration profiles of sulfate labeled degradation fragments released
In an attempt to apply gene silencing for therapeutic purposes, a by heparanase from metabolically labeled ECM in the absence
reliable system for siRNA administration, applying in vivo electro- (control) and presence of 5 g/ml and 0.25 g/ml of various hepa-
poration of siRNA expressing vectors, or lentiviral infection, was rin species are presented in Figs. (4A and B), respectively. Inhibi-
recently developed. The effectiveness of this approach for suppress- tion is reflected by the decreased amounts and Kav values of HS
ing the biological activity of heparanase was judged by the inhibi- fragments released from ECM and eluted as fractions 20-35, in
tory effect of siRNA administration in several experimental models comparison with control incubation of the ECM with recombinant
of heparanase-driven biological processes, such as delayed type heparanase in the absence of inhibitors [62, 88, 90]. Heparanase
hypersensitivity (DTH) inflammation [85], hair growth, and pros- activity is calculated as the total amount of cpm eluted in peak cor-
tate tumor progression (Vlodavsky et al., unpublished results). responding to fractions 20-35 multiplied by the Kav (i.e., elution
position) of these fragments.
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2065

Fig. (4). Representative heparanase-inhibition curves for heparins and various heparin species. [A: from Ref. 91; B: from Ref. 62]. Sulfate labeled ECM was
incubated (4 h, 37°C, pH 6.0) with recombinant human heparanase (40 ng/ml) in the absence (control) and presence of 5 g/ml (A) or 0.25 g/ml (B) heparin
or heparin derivatives. Sulfate labeled degradation fragments relesed into the incubation medum were analyzed by gel filtration on Sepharose 6B. The peak
corresponding to fractions 20-35 was used to calculate the percent residual activity of the enzyme.

As illustrated in Fig. (4A), the heparanase inhibiting activity of structure 4, corresponding to sequences of poly-pentasulfated tri-
heparin derivatives increases with increasing degrees of sulfation. saccharides separated by glycol-split uronic acid residues [94].
However, N-sulfate groups seem not essential, since their replace- The heparanase inhibitory activity (expressed as percent inhibi-
ment by N-acyl groups (N-acetyl, N-succinyl, or N-hexanoyl) tion of heparanase) of heparin and heparin derivatives at concentra-
yielded products still retaining significant inhibitory activity tion of 1 g/ml, is shown in Table 1 [62]. These data confirm that
[62,91]. (The structure of fully N-acetylated heparin (NA-H) is heparin is a strong inhibitor of heparanase (~70% inhibition at 1
represented by formula 3). As illustrated in Fig. (4B), at lower con- g/ml). Significant differences in heparanase-inhibiting activity
centrations of the inhibitor significant differences in heparanase- were associated with specific chemical modifications of heparin.
inhibition activities between heparin and N-acetyl heparin are Whereas either 6-O-desulfation or 2-O-desulfation with retention of
clearly observable. 2-O-desulfated derivatives were shown to retain L-IdoA configuration had little or no effect on the heparanase in-
the heparanase inhibition activity of the parent heparin [92]. In fact, hibitory activity of heparin, 2-O-desulfation with change of con-
neither 2-O-sulfate nor 6-O-sulfate groups are essential for efficient figuration of the L-IdoA residues to L-GalA (involved in alkali
inhibition of heparanase, provided that at least one of these posi- treatment of heparin under well defined conditions) [99]) markedly
tions is extensively sulfated [62]. Relationships between structure decreased the inhibitory activity of heparin. Also, complete removal
and heparanase inhibition activity of heparin were established using of N-sulfate groups followed by N-acetylation resulted in a substan-
a number of heparins and heparin derivatives, including some with tial decrease in the inhibitory activity (Fig. 4B). However, this ef-
various degrees of 6-O-sulfation of GlcN and 2-O-sulfation of IdoA fect was only noted for N-acetylation degrees higher than approxi-
residues, as well as “glycol-split” derivatives obtained by controlled mately 50%. On the other hand, glycol-splitting markedly increased
periodate oxidation/borohydride reduction of natural [93] or par- the heparanase-inhibiting activity of both heparins and N-acetylated
tially 2-O-desulfated heparins [94,95]. Glycol-splitting of C2-C3 heparins and restored the inhibitory effect lost upon N-acetylation
bonds of nonsulfated uronic acid residues was suggested to modu- of heparin. This effect is illustrated by data in Table 1 for N-
late the biological interactions of heparin by providing flexible acetylated heparins of the RO-type (i.e., 25% glycol-split), which
joints between protein binding sequences [62, 94-98]. Notably, almost completely inhibited the heparanase activity (to less than
when framing heparin sequences that bind FGF-2, glycol-split resi- 10% of the control at 1 g/ml and to 20-30% at 0.20-0.25 g/ml),
dues do not impair the binding to the growth factor; however, they irrespective of their degree of N-acetylation. Glycol-splitting ex-
prevent activation of FGF-2 and FGF-2-induced angiogenic activity tended to newly-generated non-sulfated IdoA/GalA residues in
[94,95]. heparin and N-acetylated heparins yielded products all showing
Heparins with various degrees of N-acetylation/N-sulfation, to- high heparanase inhibitory activity without apparent dependence on
gether with some of their glycol-split derivatives, were investigated the degree of glycol splitting. The heparanase IC50 values calculated
for their FGF-2 and heparanase-inhibition activity [62]. Partially N- from dose-dependence curves were >5 g/ml for NAH, ~0.4 g/ml
acetylated heparins listed in Table 1 had different percentages of N- for H (heparin), and ~0.2 g/ml for 100NA, RO-H [62]. As shown
acetyl groups, the complement to 100% N-substitution being N-SO3 by data in Table 1 and illustrated in Fig. (4), 100NA,RO-H emerged
groups. Glycol-split heparins and glycol-split N-acetyl heparins as an inhibitor of heparanase stronger than heparin. Data in Table 1
were of two different types. The first type, referred to as “reduced also show that it is consistently stronger than maltohexaose sulfate
oxyheparins” (RO-H) and “N-acetyl-reduced oxyheparins” (NA- (MHS, formula 8). Gel permeation chromatographic analysis of
RO-H), involved glycol-splitting of only the nonsulfated uronic some products of heparanase digestion, performed under conditions
acid residues (GlcA and IdoA) originally present in the parent hepa- of the enzyme inhibition assay, indicated that whereas heparin and
rin [62]. De-O-sulfation of one out of two IdoA residues followed N-acetyl heparin are cleaved by heparanase (as previously shown
by glycol-splitting gave heparins denoted H50gs with prevalent for heparin) [14], their glycol-split derivatives are not susceptible to
cleavage [62].
2066 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

Table 1. Distribution of Major Sulfate Groups and Heparanase Inhibitory Activity of Heparins and Derivatives [62]

Heparanase inhibitory activity, %


SO3 groups (%)
Mean [SD (N) ]a

NS A6OS I2OS 1 g/ml

HEPARINS

H1 89 79 69 73.5 [13.5 (7)]

6-O-DESULFATED H-1

H, 716OdeS (A) 81 29 55 50.5 [25.6 (7)]


77
H, 6Odes (B) 87 23 67 66.0 [ 9.4 (2)]
73
H, 6OdeS (B) 78 27 64 70.9 [18.0 (3 )]
46
H, 6OdeS (B) 82 56 68 63.4 [17.0 (6)]

2-O-DESULFATED H-1

H,IdoA (A) 83 85 0 64.7 [17.8 (3)]

H,GalA (B) 86 74 0 11.5 [11.0 (2)]

N-DESULFATED, N-ACETYLATED H-1


29
NAH 71 80 72 88.4 [11.6 (4)]
39
NAH 61 80 71 76.6 [7.4 (2)]
50
NAH 50 79 70 52.0 [ 4.4 (2)]
58
NAH 41 79 68 32.9 [15.3 (2)]
70
NAH 30 75 65 37.9 [20.2 (5)]
99
NAH 8 71 73 32.7 [18.0 (4)]
100
NAH 0 78 66 32.9 [15.3 (2)]

GLYCOL-SPLIT H-1

RO-H 89 75 67 91.3 [5.8 (2)]


52
H, gs 89 75 48 79.1 [17.7 (6)]

N-ACETYLATED, GLYCOL-SPLIT H-1


26
NA,RO-H 74 80 77 91.1 [3.1 (4)]
40
NA,RO-H 60 80 71
53
NA,RO-H 47 79 71 94.8 [3.1 (6)]
67
NA,RO-H 33 79 79 93.9 [2.2 (3)]
100
NA,RO-H 0 71 75 92.5 [5.0 (3)]
29 60
NAH, gs 71 75 40 79.6 [13.0 (2)]
43 60
NAH, gs 57 75 40 70.1 [17.2 (2)]
57 64
NAH, gs 42 75 36 87.0 [10.5 (2)]
70 59
NAH, gs 30 75 41 87.8 [12.5 (2)]

LMW HEPARINS and DERIVATIVES H-1

LMW-H-1 6.5KD 82 77 66 47.7 [35.0 (2)]


49
H, gs 6.3 KD 87 75 51 86.2 [ 2.8 (3)]
49
H, gs 3.0KD 89 75 51 [10.0 (2)]
50
LMW, NA,RO-H 5.4KD 50 79 75 90.4 [3.0 (3)]

MHS ~2.5 KD - - - 69.2 [5.7. (4)]


a 2
SD standard deviation = ((yi-ymean) /(N-1)). N = number of experiments.
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2067

4.3. Effect of Modified Heparins on Angiogenic Properties, Re- IdoA residues) are superimposable on that of the parent heparin
lease of ECM-Bound FGF-2, and Stimulation of FGF-2 Mito- (data not shown). Data in Fig. (5) confirm that glycol-split, N- ace-
genic Activity tylated heparins behave similarly to non glycol-split NAH [62] in
Some species of heparin, and especially most of its glycol-split that they release ECM-bound FGF-2 consistently less than un-
derivatives that are potent inhibitors of heparanase, are endowed modified heparin. 100NAH (not shown) and 100NA,RO.H exhibited
with antiangiogenic properties also because they prevent activation the lowest FGF-2 releasing activity among the tested compounds,
of growth factors such as FGF-2 [94, 95] and VEGF [100]; re- yielding only about twice the spontaneous release observed in pres-
viewed in [101]. A significant relationship was found between the ence of the buffer (PBS) alone.
extent of glycol-splitting and the FGF2-antagonist/angiostatic ac- The ability of heparin, 100NAH and 100NA,RO.H to promote the
tivities of glycol-split heparins [95]. However, comparison of struc- mitogenic activity of recombinant FGF-2 was investigated using a
ture-activity relationships for inhibition of growth factors and inhi- cytokine-dependent, heparan sulfate deficient, lymphoid cells
bition of their release from HSPGs (via inhibition of heparanase) (BaF3) engineered to express FGFR1. Unlike heparin, both fully N-
underlines significant differences. Notably, whereas unmodified acetylated heparin (100NAH) and its glycol-split counterpart mole-
heparin is a potent inhibitor of heparanase [62, 87], it is either inac- cule (100NA,RO.H) failed to stimulate the mitogenic activity of
tive or even pro-angiogenic in typical tests such as the CAM [94] FGF-2, beyond the basal level obtained in the absence of the added
and several in vivo assays [R Giavazzi, C Pisano et al., unpub- heparin (Fig. 6). Thus, while glycol splitting of NAH fully restores
lished]. On the other hand, whereas glycol-split N-sulfated heparins its heparanase-inhibiting activity, it fails to induce a similar restora-
inhibit both growth factors and heparanase, their N-acetylated ana- tion of the ability to displace ECM-bound FGF-2 and to stimulate
logs are inactive towards most growth factors but are as active as the mitogenic activity of recombinant FGF-2 [62].
their N-sulfated analogs in inhibiting heparanase. Inhibition of
growth factors and inhibition of heparanase show also different
dependences on the lenght of heparin sequences framed by glycol-
split residues [62,95, and unpublished results].
Some of the heparin derivatives were tested for their capacity to
release FGF-2 from ECM [62, 87]. As illustrated in Fig. (5), dose-
response curves of the FGF-2-releasing activity of glycol-split
heparin (H,52gs) and its corresponding low-molecular weight de-
rivative (LMW-H,49gs) were almost superimposable to those re-
ported for heparin [62], indicating that glycol-splitting does not
substantially modify the FGF-2-releasing properties of heparins.
Since glycol-splitting impairs oligomerization of FGF-2 [94, 95], it
is expected that glycol-split heparins release the growth factor
largely in an inactive form. Also, the curves of RO-H (i.e., heparin
glycol-split only at the level of pre-existing nonsulfated GlcA and

Fig. (5). Effect of combined N-acetylation and glycol-splitting on release Fig. (6A). Idealized representation of stimulation by heparin of FGF-2-
of ECM-bound FGF-2. ECM-coated wells were incubated (3 h, 24°C) with induced lymphoid cell proliferation. B) Effect of N-acetylation and glycol-
iodinated FGF-2 and the unbound FGF-2 was washed away. The ECM was splitting on stimulation of the mitogenic activity of FGF-2. BaF3 (F32)
then incubated (3 h, 24°C) with the indicated species of N-acetyl and glycol- lymphoid cells were plated into 69-well microtiter plates in the presence of
split heparins and aliquots of the incubation medium were counted in a 2.5 ng/ml FGF-2 and increasing concentrations of heparin /H), totally N-
gamma counter. The remaining ECM was solubilized and its radioactivity acetylated heparin (NAH), or the corresponding glycol-split derivtive of
counted and used to calculate the percentage of ECM-bound 125I-FGF-2 RO-type. After 48 h, 3H-thymidine was added (1 Ci/well) and the amount
released by each compound [62]. of incorporated tymidine was determined as described in Ref. [67].
2068 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

5. MODELING CONSIDERATIONS sequences for binding to both heparin/HS binding domains 1 and 2
Determination of the molecular conformation of HS/heparin should correspond to an octa- decasaccharide, including the GlcA
chains recognized and cleaved by heparanase, as well as those that residue. Heparin tetra and hexasaccharides are poor inhibitors of
effectively inhibit its enzymatic activity, is a formidable task. Even heparanase [87, and our own unpublished data]. Most probably,
modeling based on established criteria of conformational energies these short non-GlcA-containing oligosaccharides bind to either
must take into account the extraordinary conformational versatility HBD-1 or HBD-2 (or, at high concentrations, to both), but such a
of GAG chains that contain IdoA residues. IdoA residues have been binding is weak when it is not strenghtened by any interaction with
demonstrated to be endowed with a unique conformational flexibil- the active site of the enzyme. On the other hand, small heparin
ity (“plasticity”). Such a plasticity, associated with different oligo-saccharide chains terminating with an aza-sugar (exemplified
equienergetic conformations of IdoA residues, all co-existing in a by formula 7, where R may be another heparin oligosaccharide)
rapid dynamic equilibrium, can currently explain the better protein effectively inhibit heparanase [107], suggesting that combination of
binding capacity and associated biological properties of IdoA- blocking the active site of the enzyme with the aza sugar residue
containing sequences, as compared to the more rigid GlcA- and either HBD-1 or HBD-2 (or both basic domains) with heparin
containing ones [89, 102]. As demostrated for complexes with other oligosaccharide sequences, is effective in preventing recognition
proteins such as antithrombin both in crystals [103] and in solution and cleavage of HS chains by heparanase.
[104] and for FGF-2 in crystals [105], heparin oligosaccharides Full length heparins are strong inhibitors of heparanase in spite
may select one of the most favored conformations of its IdoA (or of being also substrates for the enzyme [14]. However, not all the
IdoA2SO3) residues for optimal docking to the protein [89,106]. It SO3 groups of heparin are essential for the inhibition of the enzy-
is thus possible that some of the iduronate residues along the same matic activity. Such a behavior is compatible with an accepted
heparin (or heparin derivative) chain adopt different local confor- model of heparin chains where clusters of three sulfate groups al-
mations depending on the 3D structure of the protein to which it ternate on both sides of a helix, irrespective of the conformation of
binds. iduronate residues [108] (Fig. 8a). The assumption that only sulfate
Although fine details of the 3D structure of heparanase are not groups on one side of the helix are involved in the interaction with
yet known, examination of models thus far available permits to basic clusters of heparanase is compatible with the observation that
envisage how the enzyme binds to and cleaves the HS chains of heparins whose N-SO3 groups have been replaced up to 50% by N-
HSPGs. As illustrated in Fig. (7), an homology model of hepara- acetyl groups retain the heparanase-inhibition activity of the parent
nase using the X-ray structure of 1,4--xylanase [39] suggests that heparin [62,98]. On the other hand, complete removal of either 6-O-
HS chains can be accommodated in a canyon in the active form of sulfate or 2-O-sulfate groups does not significantly impair the
the protein [26]. The glycosidic linkage of the GlcA residue suscep- heparanase inhibition activity of heparin, provided the GAG is fully
tible to cleavage by heparanase is expected to be docked to the N-sulfated and 2-O- (or 6-O)- sulfated, respectively [62].
active site of the enzyme, which is approximately located at the The impressive effectiveness of glycol-split heparins as hepara-
center of the canyon. The oligosaccharide sequences from both nase inhibitors raises the question of the role played by glycol-split
sides of this GlcA residue conceivably bind from one side to HBD- residues in boosting such an activity. It was predicted [96] and ex-
1 and from the other side to HBD-2. Since the distance between perimentally proved [62,109] that glycol-splitting provides extra-
HBD-1 and HBD-2 is of the order of an octa-decasaccharide [M. flexibility to heparin (GAG) chains while not intefering with the
Guerrini, personal communication], the optimal length of the HS binding properties of chains segments that do not contain glycol-
split residues. As mentioned before and illustrated in Fig. (8a),
short segments of N-sulfated heparin chains can be represented as
linearly propagating helices featuring arrays of NSO3, 2SO3, and
6SO3 groups alternating on each side of the chain [108]. Because of
limited rotation allowed around ”normal” glycosidic bonds, these
arrays cannot align themselves on the same side of the heparin
chain. On the contrary, glycol-splitting introduces extra degrees of
rotational freedom and may generate novel conformations such as
those shown in Fig. (8b) and (8c). Conformation b, which is one of
the two energetically favored conformations compatible with NMR
data, features a “kink” at the level of the glycol-split residue; a sec-
ond permitted conformation compatible with experimental data,
also showing the kink, is symmetrical to the one in b [95]. It must
be noted that glycol-split heparins may also adopt conformations
such as that shown in Fig. (8c), with arrays of up to four sulfate
groups in a row. At least in principle, such conformations may be-
came favored by complexing with heparanase. More refined model-
ing studies are needed to elucidate the actual conformation adopted
by HS/heparin chains both in the absence and in the presence of
heparanase. Further studies are also needed to rationalize the nota-
ble finding that whereas replacement with N-acetyl groups of all the
N-sulfate groups of heparin impairs the heparanase inhibition activ-
ity, glycol-splitting restores this activity to the level of the corre-
sponding glycol-split, N-sulfated heparins [62]. It is most probable
that glycol-split N-acetylated heparins adopt, in their complexes
Fig. (7). Homology modeling of heparanase using the X-ray structure of with heparanase, a conformation that is different from that of N-
1,4--xylanase as a template [39].The yellow letters represent the conserved sulfated heparins, conceivably also because of requirements for
critical residues of Glu225 (proton donor) and Glu343 (nucleophile), respec- some of their N-acetyl groups to settle in hydrophobic niches of the
tively. The HS chain was manually located on the active site of heparanase, protein. Molecular modeling studies indicate that the overall con-
and the complex model was constructed using molecular mechanics and formation of N-acetyl heparin (Fig. 8d) does not substantially differ
molecular dynamics calculations with consideration of the mechanism of from that of N-sulfated heparin (Fig. 8a) [110]. Also the favored
enzymatic glycoside hydrolysis [26]. “kinked” conformations of glycol-split N-acetyl heparin (such as
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2069

that represented in Fig. 8e) do not substantially differ from the cor- residues of MHS are characterized by some flexibility associated
responding ones of the glycol-split N-sulfated heparin compatible with rotation of residues around the axial glycosidic bonds. It is
with experimental NOE values [98]. Both N-acetyl heparin in its accordingly expected that the MHS molecules (Formula 8) adjust
energetically favored conformation 8d [110] and in its “kinked” themselves in the heparanase molecule in such a way that most of
form (8e) feature only two sulfate groups (as opposed to three for the basic residues of sites HBD1 and HBD2 are blocked by the
N-sulfated heparin) regularly alternating on both sides of the GAG sulfated oligosaccharide. The recently reported heparanase inhibitor
chains. However, similarly to glycol-split N-sulfated heparins, in oligomannurarate sulfate (JG3) is suggested to bind to HBD1
some of the energetically allowed conformations of glycol-split N- [111].
acetyl heparins such as that shown in Fig. (8f), arrays of sulfate
groups of adjacent disaccharide units may be brought close to each 6. CONCLUSIONS
other on the same side of the GAG chain [S. Guglieri et al., unpub- Although a significant progress has been made during the last
lished results]. Heparin/HS chains in these conformations may be years in understanding heparanase biology, there is much to be
stabilized and selected by heparanase to form stable complexes. learned. Accumulation of compelling evidence implies that the
Thus, the combination of local flexibility associated with the intrin- enzyme is up regulated in primary human tumors and inversely
sic plasticity of sulfated iduronate residues and the extra-flexibility correlates with survival rate of cancer patients post operation. An-
induced by glycol-splitting seems to favor optimal “exploitation” of giogenesis is the primary 'suspect' that governs heparanase medi-
sulfate groups in binding to basic sites lying along the canyon of ated tumor progression. This mode of action and the related clinical
heparanase and lead to effective inhibition of the enzyme by hepa- applications await further confirmation and require new molecular
rin derivatives having not more than two sulfate groups per disac- tools such as small inhibitory molecules, neutralizing antibodies,
charide unit. and heparanase-inhibiting glycol-split species of heparin with in
Although not all the sulfate groups of a extensively sulfated vivo pharmacological and pharmacokinetics tailored to the specific
polysaccharide are necessarily involved in protein binding [3], therapeutic application, as well as an effective approach to deliver
some persulfated poly- and oligosaccharides are often stronger anti-heparanase agents into the tumor tissue. The ability of hepara-
binders than the corresponding less sulfated species [86, and our nase to function in an apparently enzymatic independent manner,
own unpublished data]. Data in Table 1 indicate that the hexa- noted in several experimental settings [55, 65, 70, 72, 73], is in-
saccharide MHS (which, as shown in formula 8, may bear up to triguing and affects the way the protein is envisioned. Thus, while
three sulfate groups per each internal disaccharide residue and up to attention was mainly focused on compounds that inhibit heparanase
four for the terminal ones), though less potent than glycol-split enzymatic activity, no information is available on protein domains
heparins, is also a good inhibitor of heparanase [86]. Carbohydrate responsible for the non-enzymatic functions of the heparanase en-
backbones such as those generated by the -1,4-linked glucose zyme. In this respect, identification of a putative heparanase recep-

Fig. (8). Minimized structure of heparin octasaccharide sequences. (a) heparin; (b) and (c), glycol-split heparin in its kinked and extended form, respec-
tively; (d) N-acetylated heparin; (e) and (f), glycol-split N-acetylated heparin in its kinked and extended form, respectively. (See text). Model a) was con-
structed using pdb code 1HPN [108], with IdoA2SO3 residues in the 1C4 conformation. Minimized structures b-f were obtained by modifying a). Arrows indi-
cate IdoA2SO3 residues that are 2-O-desulfated and glycol-split in strucures b-f. Glycol-split residues are circled.
2070 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

tor is a major future challenge. Some of the N-acetylated glycol- GPI-anchored = Glycosyl phosphatidyl inositol-anchored
split heparins (such as 100NA, ROH) emerge as promising hepara- TNF = Tissue necrosis factor 
nase-inhibiting lead compounds. Another important objective is the
establishment of a reliable diagnostic assay to monitor heparanase PKC = Protein kinase C
levels and activity in plasma and urine of cancer patients by mean PKA = Protein kinase A
of a sensitive, high throughput activity assay and ELISA method. HDB = Heparin binding domain
The secreted nature of the enzyme and its induction in primary
RLP = Lipoprotein receptor-related protein
human tumors predict that under certain conditions the protein is
present in body fluids. It is conceivable that elevated levels of RT-PCR = Reverse transcriptase polymerase chain
heparanase are found in the plasma and/or urine of cancer patients reaction
as well as in other pathological disorders such as diabetes [112- MRI = Magnetic resonance imaging
114]. Induction of heparanase already at early phases of cancer SMA = Smooth muscle actin
progression supports an important diagnostic and, possibly, prog-
nostic value of such assays. HB-EFG = Heparin-binding epidermal growth factor
Recent publications clearly imply that heparanase may be in- KGF = Keratinocyte growth factor
volved in pathological conditions other than cancer. An intriguing Hpa-tg = Heparanase transgenic mice
example is the observation that heparanase over expressing trans- PECAM-1 = Pletelet endothelial cell adhesion molecule-1
genic mice escaped amyloid deposition in experimental models of
inflammatory-associated amyloidosis [115] and prion diseases (our HBP = Heparin binding protein
unpublished observations). The possible involvement of heparanase Cyto.D = Cytochalasin D
in degenerative diseases such as Alzheimer's disease is only starting MRS = 2'-Deoxy-N6-methyladenosine-3',5'-
to emerge and should be perused, taking into account the increasing bisphosphate
significance of these illnesses. Likewise, the causal involvement of
PPADS = Pyridoxal-phosphate-6-azophenyl-2'-
heparanase in diabetic nephropathy [116], auto-immunity [117] and
4'disulphonic acid
inflammatory disorders [85] should be investigated.
siRNA = Small interference RNA
Functional domains other than the basic heterodimer structure
[35-37] and amino acids (Glu225, Glu343) critical for the enzyme DTH = Delayed type hypersensitivity
catalytic activity [22], have not been so far identified in the hepara- GlcA = Glucuronic acid
nase protein, making screens for inhibitory molecules random in MHS = Maltohexaise sulfate
nature. The identification of heparin binding domains [56] and the
ability of the corresponding KKDC peptide, especially when ap- PI-88 = Phosphomannopentaose sulfate
plied as a dimer, to inhibit heparanase enzymatic activity [56], GAG = Glycosaminoglycan
clearly emphasize the need for crystallization and accurate under- AT = Antithrombin
standing of the 3D structure of the enzyme toward an efficient drug
development program. While it is now well accepted that a single GFR = Growth factors
active heparanase enzyme is expressed by mammals, heparanase NAH = N-acetyl heparin
splice variants have recently been characterized in the Mole rat NA-ROH = N-acetyl reduced oxyheparin
[118] and human (our unpublished results), although their role has
IdoA = Iduronic acid
not been established. The heparanase system may be envisaged to
include heparanase 1 and its splice variants, the three splice variants GalA = Galacturonic acid
of heparanase 2 [119], the heparanase processing protease, and, CAM = Chicken chorioallantoic membrane
possibly, the heparanase cell surface receptor. Gain of more infor- LMW-H = Low-molecular weight heparin
mation will enable the development of new inhibitory strategies
directed against the enzymatic and non-enzymatic functions of RO.H = Reduced oxyheparin
heparanase, altogether offering better therapeutic opportunities. FGFR1 = Fibroblast growth factor receptor 1
NOE = Nuclear Overhauser effect.
ACKNOWLEDGMENTS
Original work from our groups summarized and discussed in REFERENCES
this review was supported by grants from The Israel Science Foun- [1] Kjellén L, Lindahl U. Proteoglycans: structures and interactions.
dation (grant 549/06); National Cancer Institute, NIH (grant RO1- Annu Rev Biochem 1991; 60: 443-75.
CA106456); The European Union Project HEPARANASE (Con- [2] Iozzo RV, San Antonio JD. Heparan sulfate proteoglycans: heavy
tract QLK3-CT-2002-02049); The Rappaport Family Institute hitters in the angiogenesis arena. J Clin Invest 2001; 108: 349-55.
Fund, and Sigma-Tau, Pomezia, Italy. We gratefully acknowledge [3] Casu B, Lindahl U. Structure and biological interactions of heparin
the contribution, motivation and assistance of the research teams in and heparan sulfate. Adv Carboh Chem Biochem 2001; 57: 159-
the Tumor Biology Research Unit of the Hadassah-Hebrew Univer- 206.
sity Medical Center (Jerusalem, Israel) and the Cancer and Vascular [4] Bernfield M, Gotte M, Park PW, Reizes O, Fitzgerald ML,
Biology Research Center of the Rappaport Faculty of Medicine Lincecum J, et al. Functions of cell surface heparan sulfate proteo-
(Technion, Haifa). Molecular modeling of heparin-derived oligo- glycans. Annu Rev Biochem 1999; 68: 729-77.
saccharides by Marco Guerrini and Sara Guglieri (Ronzoni Insti- [5] Whitelock JM, Iozzo R. Heparan sulfate: a complex polymer
tute) is also gratefully aknowledged. charged with biological activity. Chem Rev 2005; 105: 2745-64.
[6] Vlodavsky I, Folkman J, Sullivan R, Friedman R, Ishai-Michaeli
ABBREVIATIONS R, Sasse J, et al. Endothelial cell-derived basic fibroblast growth
HS = Heparan sulfate factor: synthesis and deposition into subendothelial extracellular
matrix. Proc Natl Acad Sci USA 1987; 84: 2292-6.
ECM = Extracellular matrix [7] Folkman J, Klagsbrun M, Sasse J, Wadzinski M, Ingber D, Vlo-
HSPG = Heparan sulfate proteoglycan davsky I. A heparin-binding angiogenic protein--basic fibroblast
FGF-2 = Fibroblast growth factor 2 growth factor-- is stored within basement membrane. Am J Pathol
1988; 130: 393-400.
VEGF = Vascular endotelial growth factor
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2071

[8] Park JE, Keller GA, Ferrara N. The vascular endothelial growth [31] Klerk CP, Smorenburg SM, Otten HM, Lensing AW, Prins MH,
factor (VEGF) isoforms: differential deposition into the subepithe- Piovella F, et al. The effect of low molecular weight heparin on
lial extracellular matrix and bioactivity of extracellular matrix- survival in patients with advanced malignancy. J Clin Oncol 2005;
bound VEGF. Mol Biol Cell 1993; 4: 1317-26. 23: 2130-5.
[9] Garg HG, Linhardt RJ, Hales CA Eds. Chemistry and biology of [32] Lee AY, Rickles FR, Julian JA, Gent M, Baker RI, Bowden C, et
heparin and heparan sulfate. Elsevier, Amsterdam 2005. al. Randomized comparison of low molecular weight heparin and
[10] Timpl R, Brown JC. Supramolecular assembly of basement mem- coumarin derivatives on the survival of patients with cancer and
branes. Bioessays 1996; 18: 123-32. venous thromboembolism. J Clin Onc 2005; 23 : 2123-9.
[11] Stetler-Stevenson WG, Aznavoorian S, Liotta LA. Tumor cell [33] Rickles FR. If heparanase is the answer, what is the question? J
interactions with the extracellular matrix during invasion and me- Thromb Haemost 2006; 4: 557-9.
tastasis. Annu Rev Cell Biol 1993; 9: 541-73. [34] Goldshmidt O, Zcharia E, Aingorn H, Guatta-Rangini Z, Atzmon
[12] Werb Z. ECM and cell surface proteolysis: regulating cellular R, Michal I, et al. Expression pattern and secretion of human and
ecology. Cell 1997; 91: 439-42. chicken heparanase are determined by their signal peptide se-
[13] Freeman C, Parish CR. Human platelet heparanase: purification, quence. J Biol Chem 2001; 276: 29178-87.
characterization and catalytic activity. Biochem J 1998; 330:1341- [35] Fairbanks MB, Mildner AM, Leone JW, Cavey GS, Mathews WR,
50. Drong RF, et al. Processing of the human heparanase precursor
[14] Sandbäck-Pikas D, Li J-p, Vlodavsky I, Lindahl U. Substrate and evidence that the active enzyme is a heterodimer. J Biol Chem
specificity of heparanases from human hepatoma and platelets. J 1999; 274: 29587-90.
Biol Chem 1998; 273: 18770-7. [36] Levy-Adam F, Miao HQ, Heinrikson RL, Vlodavsky I, Ilan N.
[15] Vlodavsky I, Goldshmidt O. Properties and function of heparanase Heterodimer formation is essential for heparanase enzymatic activ-
in cancer metastasis and angiogenesis. Haemostasis 2001; 31: 60- ity. Biochem Biophys Res Commun 2003; 308: 885-91.
3. [37] McKenzie E, Young K, Hircock M, Bennett J, Bhaman M, Felix
[16] Okada Y, Yamada S, Toyoshima M, Dong J, Nakajima M, R, et al. Biochemical characterization of the active heterodimer
Sugahara K. Structural recognition by recombinant human hepara- form of human heparanase (Hpa1) protein expressed in insect
nase that plays critical roles in tumor metastasis. J Biol Chem cells. Biochem J 2003; 373: 423-35.
2002; 277: 42488-95. [38] Simizu S, Ishida K, Wierzba MK, Osada H. Secretion of hepara-
[17] Gong F, Jemth P, Escobar Galvis ML, Vlodavsky I, Horner A, nase protein is regulated by glycosylation in human tumor cell
Lindahl U, et al. Processing of macromolecular heparin by hepara- lines. J Biol Chem 2004; 279: 2697-703.
nase. J Biol Chem 2003; 278: 35152-8. [39] Hulett MD, Hornby JR, Ohms SJ, Zuegg J, Freeman C, Gread J, et
[18] Bame KJ. In: Garg HG, Linhardt RJ, Hales CA Eds, Heparan al. Identification of active-site residues of the pro-metastatic en-
sulfate degradation by heparanases. Elsevier, Amsterdam 2005; doglycosidase heparanase. Biochemistry 2000; 39: 15659-67.
259-83. [40] Shteper PJ, Zcharia E, Ashhab Y, Peretz T, Vlodavsky I, Ben-
[19] Parish CR, Freeman C, Hulett MD. Heparanase: a key enzyme Yehuda D. Role of promoter methylation in regulation of the
involved in cell invasion. Biochim Biophys Acta 2001; 1471: mammalian heparanase gene. Oncogene 2003; 22: 7737-49.
M99-108. [41] Ogishima T, Shiina H, Breault JE, Terashima M, Honda S,
[20] Vlodavsky I, Friedmann Y. Molecular properties and involvement Enokida H, et al. Promoter CpG hypomethylation and transcription
of heparanase in cancer metastasis and angiogenesis. J Clin Invest factor EGR1 hyperactivate heparanase expression in bladder can-
2001; 108: 341-7. cer. Oncogene 2005; 24: 6765-72.
[21] Dempsey LA, Brunn GJ, Platt JL. Heparanase, a potential regula- [42] de Mestre AM, Rao S, Hornby JR, Soe-Htwe T, Khachigian LM,
tor of cell-matrix interactions. Trends Biochem Sci 2000; 25: 349- Hulett MD. Early growth response gene 1 (EGR1) regulates
51. heparanase gene transcription in tumor cells. J Biol Chem 2005;
[22] Hulett MD, Freeman C, Hamdorf BJ, Baker RT, Harris MJ, Parish 280: 35136-47.
CR. Cloning of mammalian heparanase, an important enzyme in [43] Elkin M, Cohen I, Zcharia E, Orgel A, Guatta-Rangini Z, Peretz T,
tumor invasion and metastasis. Nat Med 1999; 5: 803-9. et al. Regulation of heparanase gene expression by estrogen in
[23] Kussie PH, Hulmes JD, Ludwig DL, Patel S, Navarro EC, Seddon breast cancer. Cancer Res 2003; 63: 8821-6.
AP, et al. Cloning and functional expression of a human hepara- [44] Baraz B, Haupt Y, Elkin M, Peretz T, Vlodavsky I. Tumor sup-
nase gene. Biochem Biophys Res Commun 1999; 261: 183-7. pressor p53 regulates heparanase gene expression. Oncogene, On-
[24] Toyoshima M, Nakajima M. Human heparanase. Purification, line, March 2006.
characterization, cloning, and expression. J Biol Chem 1999; 274: [45] Nardella C, Lahm A, Pallaoro M, Brunetti M, Vannini A, Steinku-
24153-60. hler C. Mechanism of activation of human heparanase investigated
[25] Vlodavsky I, Friedmann Y, Elkin M, Aingorn H, Atzmon R, Ishai- by protein engineering. Biochemistry 2004; 43: 1862-73.
Michaeli R, et al. Mammalian heparanase: gene cloning, expres- [46] Abboud-Jarrous G, Rangini-Guetta Z, Aingorn H, Atzmon R,
sion and function in tumor progression and metastasis. Nat Med Elgavish S, Peretz T, et al. Site-directed mutagenesis, proteolytic
1999; 5: 793-802. cleavage, and activation of human proheparanase. J Biol Chem
[26] Simizu S, Ishida K, Osada H. Heparanase as a molecular target of 2005; 280: 13568-75.
cancer chemotherapy. Cancer Sci 2004; 95: 553-8. [47] Gingis-Velitski S, Zetser A, Kaplan V, Ben-Zaken O, Cohen, E,
[27] Ferro V, Hammond E, Fairweather JK. The development of inhibi- Levy-Adam F, et al. Heparanase uptake is mediated by cell mem-
tors of heparanase, a key enzyme involved in tumour metastasis, brane heparan sulfate proteoglycans. J Biol Chem 2004; 279:
angiogenesis and inflammation. Mini Rev Med Chem 2004; 4: 44084-92.
693-702. [48] Vreys V, Delande N, Zhang Z, Coomans C, Roebroek A, Durr J, et
[28] Sasisekharan R, Shriver Z, Venkataraman G, Narayanasami U. al. Cellular uptake of mammalian heparanase precursor involves
low density lipoprotein receptor-related proteins, mannose 6-
Roles of heparan-sulfate glycosaminoglycans in cancer. Nat Rev
Cancer 2002; 2: 521-8. phosphate receptors, and heparan sulfate proteoglycans. J Biol
Chem 2005; 280: 33141-8.
[29] Liu D, Shriver Z, Qi Y, Venkataraman G, Sasisekharan R. Dy-
[49] Zetser A, Levy-Adam F, Kaplan V, Gingis-Velitski S, Bashenko
namic regulation of tumor growth and metastasis by heparan sul-
fate glycosaminoglycans. Semin Thromb Hemost 2002; 28: 67-78. Y, Schubert S, et al. Processing and activation of latent heparanase
occurs in lysosomes. J Cell Sci 2004; 117: 2249-58.
[30] Kakkar AK, Levine MN, Kadziola Z, Lemoine NR, Low V, Patel
[50] Nadav L, Eldor A, Yacoby-Zeevi O, Zamir E, Pecker I, Ilan N, et
HK, et al. Low molecular weight heparin, therapy with Dalteparin,
al. Activation, processing and trafficking of extracellular hepara-
and survival in advanced cancer: the Fragmin advanced malig-
nancy outcome study (FAMOUS). J Clin Oncol 2004; 22: 1944-8. nase by primary human fibroblasts. J Cell Sci 2002; 115: 2179-87.
2072 Current Pharmaceutical Design, 2007, Vol. 13, No. 20 Casu et al.

[51] Goldshmidt O, Nadav L, Aingorn H, Irit C, Feinstein N, Ilan N, et [71] Joyce JA, Freeman C, Meyer-Morse N, Parish CR, Hanahan DA.
al. Human heparanase is localized within lysosomes in a stable Functional heparan sulfate mimetic implicates both heparanase and
form. Exp Cell Res 2002; 281: 50-62. heparan sulfate in tumor angiogenesis and invasion in a mouse
[52] Cohen E, Atzmon R, Vlodavsky I, Ilan N. Heparanase processing model of multistage cancer. Oncogene 2005; 24: 4037-51.
by lysosomal/endosomal protein preparation. FEBS Lett 2005; [72] Goldshmidt O, Zcharia E, Cohen M, Aingorn H, Cohen I, Nadav
579: 2334-8 L, et al. Heparanase mediates cell adhesion independent of its en-
[53] Fuller M, Chau A, Nowak RC, Hopwood JJ, Meikle PJA. Defect zymatic activity. FASEB J 2003; 17: 1015-25.
in exo-degradative pathways provides insight into endo- [73] Gingis-Velitski S, Zetser A, Flugelman MY, Vlodavsky I, Ilan N.
degradation of heparan and dermatan sulfates. Glycobiology 2005; Heparanase induces endothelial cell migration via protein kinase
16: 318-25. B/Akt activation. J Biol Chem 2004; 279: 23536-41.
[54] Chen G, Wang D, Vikramadithyan R, Yagyu H, Saxena U, Pillar- [74] Ellis LM, Staley CA, Liu W, Fleming RY, Parikh NU, Bucana
isetti S, et al. Inflammatory cytokines and fatty acids regulate en- CD, et al. Down-regulation of vascular endothelial growth factor
dothelial cell heparanase expression. Biochemistry 2004; 43: 4971- in a human colon carcinoma cell line transfected with an antisense
7. expression vector specific for c-src. J Biol Chem 1998; 273: 1052-
[55] Sotnikov I, Hershkoviz R, Grabovsky V, Ilan N, Cahalon L, Vlo- 7.
davsky I, et al. Enzymatically quiescent heparanase augments T [75] Jiang P, Kumar A, Parrillo JE, Dempsey LA, Platt JL, Prinz RA, et
cell interactions with VCAM-1 and extracellular matrix compo- al. Cloning and characterization of the humane heparanase-1
nents under versatile dynamic contexts. J Immunol 2004; 172: (HPR-1) gene promoter: role of GA-binding protein and Sp! In
5185-93. regulating HPR1 basal promoter activity. J Biol Chem 2002; 277:
[56] Levy-Adam F, Abboud-Jarrous G, Guerrini M, Beccati D, Vlo- 8989-98.
davsky I, Ilan N. Identification and characterization of hepa- [76] Mukhopadhyay D, Tsiokas L, Zhou XM, Foster D, Brugge JS,
rin/heparan sulfate binding domains of the endoglycosidase Sukhatme VP. Hypoxic induction of human vascular endothelial
heparanase. J Biol Chem 2005; 280: 20457-66. growth factor expression through c-Src activation. Nature 1995;
[57] Elkin M, Ilan N, Ishai-Michaeli R, Friedmann Y, Papo O, Pecker I, 375: 577-81.
et al. Heparanase as mediator of angiogenesis: mode of action. [77] Nakajima M, Irimura T, Di Ferrante N, Nicolson GL. Metastatic
FASEB J 2001; 15: 1661-3. melanoma cell heparanase. Characterization of heparan sulfate de-
[58] Friedmann Y, Vlodavsky I, Aingorn H, Aviv A, Peretz T, Pecker I, gradation fragments produced by B16 melanoma endoglucuronida-
et al. Expression of heparanase in normal, dysplastic, and neoplas- se. J Biol Chem 1984; 259: 2283-90.
tic human colonic mucosa and stroma. Evidence for its role in [78] Nakajima M, DeChavigny A, Johnson CE, Hamada J, Stein CA,
colonic tumorigenesis. Am J Pathol 2000; 157: 1167-75. Nicholson GL. Suramin. A potent inhibitor of melanoma hepara-
[59] Vlodavsky I, Miao HQ, Medalion B, Danagher P, Ron D. In- nase and invasion. J Biol Chem 1991; 266: 9661-6.
volvement of heparan sulfate and related molecules in sequestra- [79] Vlodavsky I, Fuchs Z, Bar-Ner M, Ariav Y, Schirrmacher V.
tion and growth promoting activity of fibroblast growth factor. Lymphoma cells mediated degradation of sulfated proteoglycans in
Cancer Metastasis Rev 1996; 15: 177-86. the subendothelial extracellular matrix: relation to tumor cell me-
[60] Vlodavsky I, Bar-Shavit R, Ishai-Michaeli R, Bashkin P, Fuks Z. tastasis. Cancer Res 1983; 43: 2704-11.
Extracellular sequestration and release of fibroblast growth factor: [80] Edovitsky E, Elkin M, Zcharia E, Peretz T, Vlodavsky I. Hepara-
a regulatory mechanism? Trends Biochem Sci 1991; 16: 268-71. nase gene silencing, tumor invasiveness, angiogenesis, and metas-
[61] Zcharia E, Zilka R, Yaar A, Yacoby-Zeevi O, Zetser A, Metzger S, tasis. J Natl Cancer Inst 2004; 96: 1219-30.
et al. Heparanase accelerates wound angiogenesis and wound heal- [81] Ohkawa T, Naomoto Y, Takaota M, Nobuhisa T, Noma K, Mokoti
ing in mouse and rat models. FASEB J 2005; 19: 211-21. T, et al. Localization of heparanase in esophaegeal cancer cells: re-
[62] Naggi A, Casu B, Perez M, Torri G, Cassinelli G, Penco S, et al. spective roles in prognosis and differentiation. Lab Invest 2004;
Modulation of the heparanase-inhibiting activity of heparin 84: 1289-304.
through selective desulfation, graded N-acetylation, and glycol [82] Takaoka M, Nahomoto Y, Ohkawa T, Uetsuka H, Shirakawa Y,
splitting. J Biol Chem 2005; 280: 12103-13. Uno F, et al. Heparanase expression correlates with invasion and
[63] Zcharia E, Metzger S, Chajek-ShaulL T, Aingorn H, Elikn M, poor prognosis in gastric cancers. Lab Invest 2003; 83: 613-22.
Friedmann Y, et al. Transgenic expression of mammalian hepara- [83] Pan W, Miao HQ, Xu YJ, Navarro EC, Tonra JR, Corcoran E,
nase uncovers physiological functions of heparan sulfate in tissue et al. 1-[4-(1H-Benzoimidazol-2-yl)-phenyl]-3-[4-(1H-benzoimida-
morphogenesis, vascularization, and feeding behavior. FASEB J zol-2-yl)-phenyl]- urea derivatives as small molecule heparanase
2004; 18: 252-63. inhibitors. Bioorg Med Chem Lett 2006;16: 409-12.
[64] Zcharia E, Philp D, Edovitsky E, Aingorn H, Metzger S, Kleinman [84] Myler HA, Lipke EA, Rice EE, West JL. Novel heparanase-
HK, et al. Heparanase regulates murine hair growth. Am J Pathol inhibiting antibody reduces neointima formation. J Biochem (To-
2005; 166: 999-1008. kyo) 2006; 139: 339-45.
[65] Zetser A, Bashenko Y, Miao H-Q, Vlodavsky I, Ilan N. Hepara- [85] Edowitsky E, Lerner I, Zcharia E, Peretz T, Vlodavsy I, Elkin M.
nase affects adhesive and tumorigenic potential of human glyoma Role of endothelial heparanase in delayed-type hypersensi-
cells. Cancer Res 2003; 63: 7733-41. tivity. Blood 2006; 107: 3609-16.
[66] Doviner V, Maly M, Kaplan V, Gemgis-Velitsky S, Ilan S, Vlo- [86] Parish CR, Freeman C, Brown KJ, Francis DJ, Cowden WB. Iden-
davsky I, et al. Spatial and temporal heparanase expression in co- tification of sulfated oligosaccharide-based inhibitors of tumor
lon mucosa throughout the adenoma-carcinoma sequence. Modern growth and metastasis using novel in vitro assays for angiogenesis
Pathol, April 2006, on line. and heparanase activity. Cancer Res 1999; 59: 3433-41.
[67] Yang Y, Macleod V, Bendre M, Huang Y, Theus AM, Miao HQ, [87] Ishai-Michaeli R, Svahn CM, Weber M, Chajek-Shaul T, Korner
et al. Heparanase promotes the spontaneous metastasis of myeloma G, Ekre H-P, et al. Importance of size and sulfation of heparin in
cells to bone. Blood 2005; 105: 1303-9. release of basic fibroblast growth factor from the vascular endothe-
[68] Cohen I, Pappo O, Elkin M, San T, Bar-Shavit R, Hazan R, et al. lium and extracellular matrix. Biochemistry 1992; 31: 2080-8.
Heparanase promotes growth, angiogenesis and survival of pri- [88] Bar-Ner M, Eldor A, Wasserman L, Matzner Y, Cohen IR, Fuks Z,
mary breast tumors. Int J Cancer 2005; 118: 1609-17. et al. Inhibition of heparanase-mediated degradation of extracellu-
[69] Kelly T, Miao HQ, Yang Y, Navarro E, Kussie P, Huang Y, et al. lar matrix heparan sulfate by non-anticoagulant heparin species.
High heparanase activity in multiple myeloma is associated with Blood 1987; 70: 551-7.
elevated microvessel density. Cancer Res 2003; 63: 8749-56. [89] Casu B. In: Garg HG, Linhardt RJ, Hales CA Eds, Structure and
[70] Zetser A, Bashenko Y, Edovitsky E, Levy-Adam F, Vlodavsky I, active domains of heparin. Elsevier, Amsterdam 2005; 1-28.
Ilan N. Heparanase induces vascular endothelial growth factor ex- [90] Vlodavsky I, Fuks Z, Bar-Ner M, Ariav Y, Schirrmacher V. Lym-
pression: correlation with p38 phosphorylation levels and Src acti- phoma cells mediated degradation of sulfated proteoglycans in the
vation Cancer Res 2006; 66, 1455-63.
Heparanase Functions and Inhibition Current Pharmaceutical Design, 2007, Vol. 13, No. 20 2073

subendothelial extracellular matrix: relation to tumor cell metasta- thrombin and with fibloblast growth factors in solution. Semin
sis. Cancer Res 1983; 43: 2704-11. Thromb Hemost 2002; 28: 325-34.
[91] Vlodavsky I., Mohsen M, Lider O, Svahn CM, Ekre HP, Vigoda [107] Petitou M, Driguez PA. Derivés d’azasucre, inhibiteurs
M, et al. Inhibition of tumor metastasis by heparanase inhibiting d’heparanases, leur procedé de preparation, les compositions en
species of heparin. Invasion Metastasis 1994; 14: 290-302. contenant et leur utilisation. French Patent FR 2 873 377–A1. Dep.
[92] Lapierre F, Holme K, Lam L, Tressler RJ, Storm N, Wee J. 23.07.2004.
Chemical modifications of heparin that diminish its anticoagulant [108] Mulloy B, Forster M, Jones C, Davies DB. NMR and molecular
but preserve its heparanase-inhibitory, angiostatic, anti-tumor and modeling.studies in the solution conformation of heparin. Biochem
anti-metastatic properties. Glycobiology 1996; 6: 355-66. J 1993; 293: 849-58.
[93] Casu B, Diamantini G, Fedeli G, Mantovani M, Oreste P, Pescador [109] Khorramian BA. Small-angle X-ray scattering of reduced oxy-
R, et al. Retention of antilipemic activity by non-anticoagulant pe- heparins. In: Industrial Polysaccharides (Stivala SS, Crescenzi, V,
riodate-oxidized heparins. Arzneim-Forsch (Drug Res) 1986; 36: Dea ICM, Eds) Gordon and Breach, New York, 1987; 339-47.
637-42. [110] Mulloy B, Forster MJ, Jones C, Drake AF, Johnson EA, Davies
[94] Casu B, Guerrini M, Naggi A, Perez M, Torri G, Ribatti D, et al. DB. The effect of variation of substitution on the solution confor-
Short heparin sequences spaced by glycol-split uronate residues mation of heparin: a spectroscopic and molecular modeling study.
are antagonists of fibroblast growth factor 2 and angiogenesis in- Carbohr Res 1994; 255: 1-26.
hibitors. Biochemistry 2002; 41: 10519-28. [111] Zhao H, Liu H, Chen Y, Li J, Hou Y, Zhang Z, et al. Oligoman-
[95] Casu B, Guerrini M, Guglieri S, Naggi A, Perez M, Torri G, et al. nurarate sulfate, a novel heparanase inhibitor simultaneously tar-
Undersulfated and glycol-split heparins endowed with antiangio- geting basic fibroblast growth factor, combats tumor angiogenesis
genic activity. J Med Chem 2004; 47: 838-48. and metastasis. Cancer Res 2006; 66: 8779-87.
[96] Casu B. Structure, shape and function of glycosaminoglycans. In: [112] Katz A, Van-Dijk DJ, Aingorn H, Erman A, Davies M, Darmon D,
Harenberg J, Heene DL, Stehle G, Schettler G Eds, New trends in et al. Involvement of human heparanase in the pathogenesis of dia-
haemostasis. Springer Verlag, Berlin 1990; 2-11. betic nephropathy. Isr Med Assoc J 2002; 4: 996-1002.
[97] Casu B, Naggi A. Antiangiogenic heparin-derived heparan sulfate [113] Levidiotis V, Freeman C, Tikellis C, Cooper ME, Power DA.
mimics. Pure Appl Chem 2003; 75: 155-64. Heparanase is involved in the pathogenesis of proteinuria as a re-
[98] Naggi A. In: Garg HG, Linhardt RJ, Hales CA Eds, Glycol- sult of glomerulonephritis. J Am Soc Nephrol 2004; 15: 68-78.
splitting as a device for modulating inhibition of growth factors [114] Shafat I, Zcharia E, Nisman B, Nadir Y, Nakhoul F, Vlodavsky I,
and heparanase by heparin and heparin derivatives. Elsevier, Am- et al. An ELISA method for the detection and quantification of
sterdam 2005; 461-81. human heparanase. Biochem Biophys Res Commun 2006; 341:
[99] Rey RN, Perlin AS. Base-catalyzed conversion of -L-iduronic 958-63.
acid 2-sulfate into a unit of -L-galacturonic acid, and related re- [115] Li J-P, Escobar ML, Gong F, Zhang X, Zcharia E, Kisilevsky R, et
actions. Carbohydr Res 1993; 200: 437-47. al. In vivo fragmentation of heparan sulfate by heparanase overex-
[100] Pisano C, Aulicino C, Vesci L, Casu B, Naggi A, Torri G, et al. pression renders mice resitant to amyloid protein A amyloidosis.
Undersulfated, low-molecular weight glycol-split heparin as an Proc Natl Acad Sci USA 2004; 102: 6473-7.
antiangiogenic VEGF antagonist. Glycobiology 2005: 15: 1C-6C. [116] Levidiotis V, Freeman C, Tikellis C, Cooper ME, Power DA.
[101] Presta M, Leali D, Stabile H, Ronca R, Camozzi, M, Coco L, et al. Heparanase inhibition reduces proteinuria in a model of acceler-
Heparin derivatives as angiogenesis inhibitors. Curr Pharm Des ated anti-glomerular basement membrane antibody disease. Neph-
2003; 9: 553-6. rology (Carlton) 2005; 10, 167-73.
[102] Casu B, Provasoli A, Petitou M, Sinaÿ P. Conformational flexibil- [117] Lider O, Baharav E, Mekori YA, Miller T, Naparstek Y,
ity, a new concept for explaining binding and biological activities Vlodavsky I, et al. Suppression of experimental autoimmune dis-
of iduronic acid-containing glycosaminoglycans. Trends Bioch Sci eases and prolongation of allograft survival by treatment of ani-
1988; 13: 221-5. mals with low doses of heparins. J Clin Invest 1989; 83: 752-6.
[103] Jin L, Abrahams JP, Skinner R, Petitou M, Pike RN, Carrell RW. [118] Nasser NJ, Nevo E, Shafat I, Ilan N, Vlodavsky I, Avivi A. Adap-
The anticoagulant activation of antithrombin by heparin. Proc Natl tive evolution of heparanase in hypoxia-tolerant Spalax: Gene
Acad Sci USA 1977; 94: 14683-8. cloning and identification of a unique splice variant. Proc Natl
[104] Hricovini M, Guerrini M, Bisio A, Torri G, Petitou M, Casu B. Acad Sci USA 2005; 102: 15161-6.
Conformation of heparin pentasaccharide bound to antithrombin [119] McKenzie E, Tyson K, Stamps A, Smith P, Turner P, Barry R, et
III. Biochem J 2001; 359: 265-72. al. Cloning and expression profiling of Hpa2, a novel mammalian
[105] Faham S, Hileman RE, Fromm JR, Linhardt RJ, Rees DC. Heparin heparanase family member. Biochem Biophys Res Commun 2000;
276: 1170-7.
structure and interactions with basic fibroblast growth factor. Sci-
ence 1996; 271: 116-20.
[106] Hricovini M, Guerrini M, Bisio A, Torri G, Naggi A, Casu B.
Active conformations of glycosaminoglycans. NMR determination
of the conformation of heparin sequences complexed with anti-

You might also like